Sunteți pe pagina 1din 24

LRE combustor design

Contents
Contents................................................................................................... 184
List of symbols......................................................................................... 185
1 Introduction ................................................................................. 186
2 Processes occurring in the combustor ...................................... 188
3 Design and sizing of chamber.................................................... 189
3.1 Injection system.......................................................................... 189
3.2 Distributor.................................................................................... 189
3.3 Injector......................................................................................... 190
4 Combustor tube.......................................................................... 197
4.1 Size and shape........................................................................... 197
4.2 Combustion modelling................................................................ 197
4.3 Combustion stability ................................................................... 200
4.4 Pressure drop due to flow acceleration..................................... 200
4.5 Catalyst bed................................................................................ 201
4.6 Chamber throat assembly.......................................................... 202
4.7 Combustor tube wall geometry.................................................. 202
4.8 Chamber materials..................................................................... 203
4.9 Chamber wall thickness based on internal pressure................ 203
4.10 Chamber mass ........................................................................... 204
4.11 Chamber service life................................................................... 204
4.12 Other chamber characteristics................................................... 205
Problems.................................................................................................. 205
References .............................................................................................. 205
For further reading................................................................................... 206
List of symbols
Roman
A Area
C
d
Discharge coefficient
D Diameter
L Length
m Mass flow
M Mach number
n Number of injector holes
O/F Oxidizer-to-fuel mass mixture ratio
p Pressure
Q Volume flow rate
r Radius
r
a
Contraction approach radius
r
u
Longitudinal throat radius
R Specific gas constant
T Temperature
v Velocity
V Volume
Greek
| Contraction half angle
Jet angle
Density
t Residence time
Pressure loss coefficient
I Vandenkerckhove parameter
Subscripts
c Chamber or contraction
con Convergent
e Expansion
f Fuel or fluid
i Injection
o Oxidizer
t Throat
Superscripts
* Characteristic parameter
1 Introduction
A LRE rocket combustor or decomposition chamber essentially is a thin-walled vented
pressure vessel in which the rocket propellant burns or decomposes to provide a hot
high-pressure gas fit for expansion in a nozzle.
Figure 1 shows a schematic of a
combustion chamber of a large
liquid hydrogen-liquid oxygen
rocket motor. It essentially
consists of an injector and dome
assembly, an igniter tube (central
in the injector and dome) and a
combustion chamber. The injector
and dome assembly is located at
the top of the chamber. The dome
manifolds the liquid oxygen and
serves as a mount form the igniter
(middle top). The fuel is directed
via the coolant manifold and the
double wall, providing
regenerative cooling to the
combustion chamber walls, and
then to the injector. In the
combustion chamber the two
flows vaporize, mix and react
creating the hot gas needed for
thrust generation. A nozzle
extension is bolted to the aft
flange of the combustion chamber
allowing for higher performance.
Figure 2 shows a typical monopropellant decomposition chamber. It uses a catalyst
bed, placed inside the chamber and contained by retainer gauzes, to further propellant
decomposition.
The monopropellant enters the thruster via
the propellant valve and is routed directly to
the injector. The injector provides a proper
distribution of the propellant over the catalyst
bed. Under the action of the catalyst, the
monopropellant decomposes thereby
generating a hot gas mixture, which exits the
chamber through the convergent/divergent
nozzle, thereby generating thrust. Cooling of
the chamber is by radiation cooling only.
The main performance requirement for a combustion or decomposition chamber is to
achieve a high combustion quality, without unduly high mass and cost of the chamber.
Characteristic data or of some specific liquid rocket engine combustion chambers are
provided in the next table.
Figure 1: Schematic of large bipropellant rocket
combustor (courtesy Boeing)
Figure 2: Monopropellant thruster schematic
Table 1: Characteristic data of some combustion chambers
Parameter L5 LE5 HM7B HM60 (Vulcain 1) ATE
Propellant MMH/NTO LH/LOX LH/LOX LH/LOX MMH/NTO
Thrust [kN] 20 103 62.2 1140 20
Core flow O/F 2.1 5.5 5.14 6.3 2.32
Mass flow [kg/s] 6.37 28.3 13.86 262.2 5.81
Chamber
pressure [bar]
10 36.8 35 110 90
Chamber
diameter [mm]
180 240
Chamber length
[mm]
L* = 840 mm 178
Contraction ratio
[-]
3.11 10
Injector type Coaxial Coaxial Coaxial Coaxial
# of injector
elements
96 208 90 516
Injector pressure
drop [bar]
15
Cooling Regenerative Regenerative Regenerative Regenerative Regenerative
Type of wall Milled channel
wall
Brazed tubes Milled channel
wall
Milled channel
wall
Milled channel
wall
# of coolant
channels
240 128 360 122
Coolant MMH Hydrogen Hydrogen Hydrogen NTO
Material Stainless steel
liner with
galvanized nickel
closure
Nickel 200 Cu alloy inner
layer with
galvanized nickel
closure
Cu alloy inner
layer with
galvanized nickel
closure
Gold coated
NARloy Z
Maximum
chamber wall
temperature [K]
750 900 770
The table provides information of 5 different chambers. The data include general
information as motor identifier, propellants used and thrust level. Then some more
specific data are provided. Most of the data will be explained in some detail in the
following text.
2 Processes occurring in the combustor
Within an LRE combustor several processes occur, including fluid injection,
vaporization, mixing (in case of bipropellants), ignition, and combustion. These
processes are more or less subsequent to each other. This allows us to distinguish
different zones in the combustor. Typically three major zones are distinguished, see
illustration:
Figure 3: Combustion zones in a LRE combustor [Sutton]
Injection/Atomization Zone
The liquid propellants are injected into the combustion chamber via an injection
system at velocities typically between 7 to 60 m/sec. When the liquid fuel and oxidizer
are injected into the chamber the individual jets are broken up into small droplets. This
region is relatively cold; however, heat transferred via radiation from the rapid
combustion region causes most of the small droplets to vaporize. At this zone
chemical reactions are occurring, but at a minimal level since the zone is relatively
cool. Also, the region is heterogeneous, with fuel and oxidizer rich regions.
Rapid Combustion Zone
In this zone chemical reactions are fast due to the increasing temperature caused by
the liberation of heat during the reaction. Any remaining droplets are vaporized and the
mixture is fairly homogeneous due to local turbulence and diffusion of gas species.
The gas expands causing the specific volume of the mixture to increase and the gas
begins to move axially with significant velocity. There is some transverse motion of the
gas as high-burning-rate regions expand towards cooler low-burning-rate regions.
Stream Tube Combustion Zone
In this region chemical reactions continue but at a reduced rate. The axial velocity of
the gas continues to increase (200 to 600 m/sec). Transverse convective flow
decreases to almost zero and the flow forms small streamlines across which turbulent
mixing is minimal.
In actuality, the boundaries of these zones are difficult to define and transition from
one zone to the next is gradual. The length of the zones is heavily influenced by
choice of propellants and the properties unique to them, the operating conditions (i.e.
mixture ratio, chamber pressure, etc.), the injector design, the chamber geometry, and
whether an catalyst is used or not. These aspects are dealt with in some more detail in
the following sections.
3 Design and sizing of chamber
Important parameters in sizing a thrust chamber include chamber volume, shape,
mass, operating pressure, materials used, etc.
The various steps in sizing are:
- Determine chamber pressure
- Select chamber shape(s) and determine size
- Select chamber material
- Dimension chamber
- Compare results and select best design
These steps are discussed in some details below.
3.1 Injection system
Figure 4 shows the injection system of specific liquid propellant rocket motor using
UH25 as fuel and NTO as oxidizer. The liquid oxidizer enters the motor on top after
which it flows through the oxidizer manifold to the cylindrical-shaped injector. The
liquid fuel first flows into the fuel manifold. From this manifold it is fed into the
combustor via the fuel injection holes.
Figure 4: Injection system of a large liquid propellant rocket motor
The main function of the injection system is to ensure a suitable flow of the liquids
allowing for smooth mixing, vaporization, ignition and chemical reaction, all at the
proper mixture ratio. To ensure proper propellant injection, the injection system
consists of a distributor and an injector. Hereafter these two components are
discussed in some detail.
3.2 Distributor
A distributor is a manifold (an arrangement of piping/tubing) designed to evenly
distribute the propellant flow over the injector orifices while (for bipropellant motors)
ensuring a perfect sealing between the oxidizer and fuel tubes, see Figure 5 (left-hand
figure).
Figure 5: Schematic of distributor
To allow an even distribution over the injector orifices, the velocity of the liquid in the
distributor must be as low as possible. Typical flow velocities in the distributor should
be well below 10-15 m/s and at the most 20% of the injection velocity, see next
section. Once the flow velocity in the distributor has been selected, the flow cross-
sectional area can be determined using the law of mass conservation:
Q A v m = =
(3-1)
Where:
m = mass flow
= specific mass of fluid
v = flow velocity
A = flow cross-sectional area
Q = flow rate through manifold
For a bipropellant system of course we have to reckon with two fluids each with its
own density. In that case, oxidizer and fuel mass flow rate can be determined from
total mass flow rate and the O/F mass ratio. Data on propellant density may be
obtained from the literature or from measurements. In most liquid cooled rocket
motors, the distributor allows for injection of fuel close to the chamber wall. This
protects the chamber wall from overheating. The reason for taking the fuel and not the
oxidizer is that the latter may react (oxidation reaction) with the metallic chamber wall
and hence leads to corrosion.
The pressure at the inlet of the distributor (inlet of thruster) is generally referred to as
the inlet pressure. Notice that because of the low flow velocity in the distributor, the
static pressure is about equal to the total pressure. This pressure must be in excess of
the chamber pressure, but not too much, as else the feed system needed to feed the
propellants into the combustor becomes too heavy. To limit any pressure loss it is
important that the manifolds are nicely shaped with a gradual transition between pipe
sections of different size, see next section.
3.3 Injector
An injector is a disk or cylinder containing many small perforations/openings/holes,
which are usually referred to as orifices. Its purpose is to cause droplet
formation/atomization and ensure even mixing and propellant distribution over the full
cross-sectional area of the combustion chamber. This improves stability of the burning
process and reduces oscillations.
Figure 6: Injector plate (photo courtesy University of Basel)
Figure 6 shows several (7) inserts in the injector plate which each contain one large
centre perforation and 4 smaller perforations in a circle about the centre perforation.
This combination is referred to as an injector element. The surface of the injector plate
facing the combustion is generally referred to as the injector face.
Injector pattern
Figures 4-6 show that the injector holes are not arbitrarily positioned on the injector.
Generally a special pattern (arrangement) is used to allow for an even filling of the
chamber, to distribute the heat loading over the full of the face plate and to allow for
face cooling. One such pattern is a concentric pattern as shown in figures 5 and 6.
Types of injector elements
The simplest form of propellant injection in to the chamber is achieved by a shower
head injector, see Figure 7 (middle). Mixing of the fuel and the oxidizer relies on
turbulence and diffusion. Sometimes a splash plate can be used to aid the
atomization. For rapid and smooth starting, it is necessary that the injector provides an
even distribution over the full cross-sectional area of the catalyst bed. The type of
injector most widely employed is the showerhead type of injector.
Another non-impinging type of injector is the spray nozzle in which conical, solid cone,
hollow cone, or other type of spray sheet can be obtained. When a liquid hydrocarbon
fuel is forced through a spray nozzle the resulting fuel droplets are easily mixed with
gaseous oxygen and the resulting mixture readily vaporized and burned. Spray
nozzles are especially attractive for the amateur builder, since several companies
manufacture them commercially for oil burners and other applications.
A third type of non-impinging injector is the coaxial element, see figure 5 (left hand
side), where a low velocity liquid stream (oxidizer) is surrounded by a high velocity
(fuel) gas jet. This type is used in many current designs of liquid hydrogen liquid
oxygen rocket engines, like the European Ariane 5 Aestus, Vinci, and Vulcain 1 and 2
rocket motors. Advantage is that the liquid hydrogen, which is also used as coolant,
can be heated to a higher temperature before injection.
Figure 7: Schematic of non-impinging types of injectors
Besides non-impinging types of injectors, there are also many rockets that use an
impinging stream type of injector. In this type of injector, the propellants are injected
through a number of separate holes in such a manner that the fuel and oxidizer
streams impinge upon each other. Impingement aids atomization of the liquids into
droplets as well as to distribution and mixing. One type is the like-on-like impinging
injector (Figure 8) where jets of the same fluid impinge, breaking the streams into
droplets. Mixing is obtained by locating the impinging streams of fuel and oxidizer near
each other so that the resulting droplets mix well. This type of injector was used in
many liquid hydrogen-liquid oxygen rocket motors, like the Ariane 4 Viking engine.
A second type of impinging injector configuration uses jets of different fluids that
impinge on each other (Figure 9). This is for example the case in most storable, bi-
propellant, reaction control system thrusters. Depending on the thrust level, one or
more multiple unlike doublet injectors are used. Below about 100 N a single doublet
type of injector suffices [Kaiser Marquardt].
Figure 9: Unlike impinging injector configuration
Compared to the non-impinging type of injectors, the impinging type offers high
combustion efficiency, but a higher heat load on the face plate. In addition, it is very
sensitive to fabrication tolerances and hence brings high cost.
Recently investigations are concentrating on swirl type of injectors that introduce a
swirl component in the injector flow. This has been shown to enhance propellant
mixing and thus improve engine performance. It are particularly swirl coaxial injectors
that show promise for the next generation of high performance staged combustion
rocket engines utilizing hydrocarbon fields.
Selection of the best type of injector configuration is usually based on experience
obtained from existing engines. In case of a newly developed injector type, a lot of
testing has to be performed including real combustion tests in a real engine to show
that the type developed is suitable.
Dirt can build up in the orifices restricting the flow of liquid. To prevent orifices from
clogging, usually a filter screen is located in each propellant feed just upstream of the
injector. Of course, the filter screen should be of a size smaller than the size of the
orifices in the injector head.
Figure 8: Like impinging superimposed injector configuration
Dimensioning and sizing of injector orifices
Important design parameters are injection velocity, the size (area), the number of
injection holes. In this section, we will show how these parameters are related and
how they determine the jet structure.
It is important that the jet breaks up into droplets. Droplet formation increases the area
of the fluid in contact with the surrounding flow and hence improves vaporization and
the subsequent combustion and or the contact area of a liquid monopropellant with a
catalyst.
The way in which a liquid jet is resolved into drops depends on the velocity on the jet.
Capillary resolution: At flow velocities in the order of m/s, droplet formation will be
due to capillary resolution. The jet shows perpendicular constriction lines at some
distance from the holes. These constrictions increase as the jet progresses and
finally cause the formation of equidistant drops.
Oscillations in the flow: At about 10 m/s, droplet formation is caused by
oscillations in the flow. The jet performs transversal oscillations which accelerate
the formation of droplets
Atomization: At flow velocities in the order of 100 m/s, the static pressure in the jet
drops below the vapor pressure of the liquid. The ensuing vaporization causes the
jet to break up into a mist immediately on leaving the hole, this is called
atomization.
Too high an injection velocity in the axial direction of the combustion chamber may
cause the propellants to leave the motor without proper combustion taking place. This
will limit the characteristic velocity to be attained.
After the selection of a suitable injection velocity, we determine the size of the holes
and their number. From the total mass flow and the O/F ratio, the total mass flow of
the fuel and oxidizer can be determined. Each usually is injected separate from the
other. Conservation of mass dictates for each:
i i i
Q n A n v A v m = = =
(3-2)
Where:
v
i
= injection velocity
A
i
= Area of single injector hole: A
i
= A/n
n = number of injector holes or injector elements
Q
i
= flow rate through single injector hole: Q
i
= Q/n
Example: Consider a 490 N bipropellant rocket motor using NTO and MMH as
propellants. Mass mixture ratio is 1.65, and vacuum specific impulse is 320 s. Total
propellant mass flow in that case is 490/(320 x 9,81) ~ 0,15 kg/s. Based on the mass
mixture ratio we find a mass flow of about 0.10 kg/s NTO and 0.05 kg/s MMH. Fluid
density is 1450 and 874 kg/m
3
respectively. Focusing on NTO, we find that with an
injector manifold velocity of 5 m/s (well below the 10-15 m/s), the flow cross-section of
the NTO manifold should be 13.8 mm
2
. For MMH follows a value of 11.4 mm
2
or about
three times the value for NTO. The respective diameters (assuming circular cross-
section) is 4.2 and 3.8 mm, respectively. For the injector orifices to achieve an injection
velocity of 30 m/s, we find that the area of the injection holes must be 6 times smaller
than the area of the manifold in case we use a single injection hole. In case we decide
for 2 injection holes, the area should be about 3 times smaller and for 6 holes 5 times
smaller.
In practice, we find that orifice diameter typically is in the range 1-3 mm, although
diameters as small as 0.08 mm can be found. The advantages of a large diameter
are:
easier to drill;
easier to align impinging elements;
unlikely to encounter combustion instability;
less contamination sensitive.
The length of an orifice is usually chosen such that the length to diameter ratio of the
orifice is in excess of 4 (L/D > 4) and preferably around 10 to allow for fully developed
flow. This minimum length to diameter ratio is necessary to prevent the occurrence of
hydraulic flip, i.e. separation
1
of the flow from the orifice wall. It reduces the mass flow
rate of propellant and causes the mis-impingement of impinging type injectors.
Detached flow can be counteracted by further increasing the pressure which causes
the flow to reattach.
Pressure drop associated with area change
An injector element can be considered as a succession of two joints of coaxial pipes of
different diameters, see illustration below. In case of an injector flow, the liquid flows
from a large manifold into the injector tube from where it is injected into the large
combustion chamber.
In case of a flow of an incompressible medium from a large vessel into a small duct
(compare the flow of water from a bottle through the neck of the bottle), we can use
Bernoullis equation:
2
1 1
2
0 0
2
1
2
1
v p v p + = +
(3-3)
For v
0
<< v
1
:
2
1
2
1
v p = A
(3-4)
In practice, some further losses e.g. due to friction, flow separation occur associated
with how well the convergent is formed, and we find:
2
2
1
v p = A
(3-5)
Here the pressure loss caused by a change in area is defined in terms of a
(dimensionless) loss coefficient , and the flow velocity v in the smaller pipe [Bejan].
For a sudden contraction (turbulent flow) the loss coefficient is:
4 3
1 5 0
/
l
s
c
A
A
.
|
|
.
|

\
|
=
(3-6)

1
An abrupt change of flow direction at the orifice entrance reduces the local static pressure up to the
saturation pressure, and cavitation bubbles appear at the location. These bubbles make the inner flow highly
turbulent, and thus jet characteristics become dependent on the flow time. The development of cavitation
bubbles can induce for liquid flow to separate from the orifice wall.
And for a sudden expansion:
2
1
|
|
.
|

\
|
=
l
s
e
A
A

(3-7)
In case of a compression, followed by an expansion, we find that the total loss factor
is:
c e
+ =
(3-8)
In the limit case where the area of the combustion chamber and the flow cross-
sectional area of the injector manifold are much larger than the cross-sectional area of
the injector hole, we find for the compression loss factor a value of 0.5 and for the
expansion loss factor a value of 1. This then would indicate a total loss factor of 1.5 for
the injector. The loss factor may be reduced by making the transition between the two
areas more gradual. In case the loss factor has been determined from calibration
measurements, see later, the injection velocity (and hence the volumetric flow rate)
can be determined based on the pressure drop measured over the injector. It follows:

p
C
p
v
d i

=
2 2 1
(3-9)
Here C
d
is the discharge coefficient, which essentially is a down rating of the area of
an orifice or nozzle due to flow separation or friction. Like the total loss factor, it is a
dimensionless parameter characteristic for the shape of the injector and strongly
related to the flow conditions in the injector. Using the earlier calculated value of 1.5,
we find a value for the discharge coefficient of about 0.82 (value of 1 means no loss).
In practice, for a well-shaped injector nozzle a value between 0.5-0.7 is feasible.
Figure 10 shows typical pressure drop over an injector with a discharge coefficient of
0.7 in relation to injection velocity of a fluid with a mass density of 1000 kg/m
3
.
Figure 10: Injector pressure drop
For small thrusters that usually operate at low feed pressure, we find that injection
velocity should be below about 40 50 m/s (depending on the liquid density) as else
the pressure drop becomes too high. For motors operating at higher pressures, higher
injection velocities are possible. For motors using hydrogen, even higher velocities are
possible, since hydrogen density is about a factor 10-15 lower than for most common
propellants.
Given the minimum pressure drop required (Huzel/Humble), we also find that there is
a minimum value for the injection velocity with regard to stability. Based on a minimum
pressure drop of 20% we find for a liquid with a density equal to water a minimum
injection velocity of about 25 m/s at a chamber pressure of 20 bar.
Jet angle
Impinging types of injectors disintegrate a jet by impact with one or more other jets.
The level of disintegration is governed by amongst others the velocity of the jets, and
the angle at which they intersect. The angle between a jet and the chamber axis is
referred to as the jet angle, . In case of two jets with respective jet angles
o
and
f
, we
find that the resulting angle
r
is determined by the respective momentum. It follows:
f f f o o o
f f f o o o
r
cos v m cos v m
sin v m sin v m
tan
+

=
(3-10)
The larger the angle between the two jets, the better the droplet formation. Typical
angles are in between 40-100 degrees. In case of like-on-like impingement, the jet
angle is identical for both jets. In case of unlike impingement, this may differ,
depending on the criterion used for the momentum of the combined jet. In most cases
we try to have zero momentum in radial direction and the jet travelling in axial
direction.
Impingement distance for doublets or triplets should be about 5 to 7 orifice diameters
in order to limit the heat loads on the injector face.
Water flow calibration
The injector discharge coefficient is usually determined experimentally. This is
sometimes referred to as calibration of the injector. It is usually performed by recording
pressure differences versus mass flow rate. The latter is determined using weighing
tanks and time recordings. Use can also be made of volumetric flow rate, in case we
have a scale on the tank. The discharge coefficient changes with changes in Reynolds
Number and is correct for only one set of conditions. This is an important point and
cannot be stressed enough. Calibration is also performed to check the flow pattern
and orifice alignment.

Ap
A C Q
i d i

=
2
(3-11)
Here Q is the volumetric flow rate.
Injector structural design loads
According to Huzel, the main loads to be considered in the structural design of the
injectors result from propellant pressures behind the injector face and in the manifolds.
During steady state operation, the pressure load on the injector face is equal to the
injector pressure drop:
i inj
p p A =
(3-12)
During start transients, however, maximum pressure loads on the injector may be
substantially higher than during steady state. When the propellant valves are opened
rapidly, severe hydraulic ram can occur. This pressure load can be estimated
empirically as:
i inj
p 4 p A =
(3-13)
Development
The development of a rocket injector is a costly business. For example, in 2003
GenCorp Aerojet was awarded a $485,000 contract to design and test a high
performance/high technology rocket injector, using MON-25/Monomethylhydrazine
propellants, for use in a Martian-simulated environment. The development will be a
seven-month effort. The initial effort of the program will focus on injector design
characteristics required to produce high performance and stable combustion using low
temperature propellants which have freezing points below -50
o
C, similar to the
Martian environment. Subsequent phases will further develop and test new-
technology lightweight components intended for the final flight version of the engine.
4 Combustor tube
4.1 Size and shape
Because of the high pressure in such combustors, pressures up to 200 bars are not
uncommon; the shape is kept very simple, being mostly of a cylindrical or spherical
nature, see illustration.
The cylindrical shape has the advantage of easy
manufacturing. The spherical yields a minimum surface
area for a given volume. Other shapes include the pear-
shape, which is found particularly in high-thrust rocket
motors, and the tubular and conical shape. The latter are
without a throat-section, which eases manufacturing.
All processes occurring in the combustion chamber
(vaporization, mixing, chemical reactions, etc.) take some
time to happen. The minimum size of the combustion
chamber while still ensuring satisfactory combustion is
determined by the time needed for vaporization, mixing,
ignition and chemical reactions:
If the chamber is too short, part of the energy will be
released outside the chamber and hence does not
contribute optimally to the thrust;
If the chamber is too long, thermal energy (heat) will
leak away to the combustor wall and hence, thrust
decreases. In addition, the thrusters will be relatively
heavy.
So, it seems like that there is an optimum size for the
combustion chamber.
4.2 Combustion modelling
A liquid propellant can either be hypergolic or non-hypergolic. Hypergolic propellants
react spontaneously when mixed in the chamber without the use of an igniter. The
various elements of the combustion process of hypergolic propellants are depicted in
Figure 12a. A combination of a diffusion
2
flame and a premixed
3
flame is possible. The
diffusion flame is caused by reaction between the oxidiser and fuel that are vaporised
and react in the gas phase. This mixture might be preceded by mixing in the liquid
phase.

2
In a diffusion flame, there fuel and oxidizer are originally separated. The mixing and combustion reactions
take place together.
3
In a premixed flame, the fuel and oxidizer are mixed before reaching the flame.
Figure 11: Geometry of combustion chamber with nozzle
Hypergolic propellants
Mixing in liquid phase
Chemical reactions in liquid
phase
Vaporisation and reactions in
vapour phase
Mixing of gas and vapours
Vaporisation and chemical
reactions in vapour phase
Atomisation of propellants
Diffusion flame,
combustion of oxidiser
and fuel droplets
Combustion in gaseous phase
Premixed flame
Combustion products
D
i
f
f
u
s
i
o
n

o
f

t
h
e

i
n
t
e
r
m
e
d
i
a
t
e

r
e
a
c
t
i
o
n

p
r
o
d
u
c
t
s

i
n

t
h
e

d
i
r
e
c
t
i
o
n

o
p
p
o
s
i
t
e

t
o

t
h
e

f
l
o
w

H
e
a
t

t
r
a
n
s
f
e
r

t
o

t
h
e

l
i
q
u
i
d

a
n
d

g
a
s
e
o
u
s

p
h
a
s
e
s

b
y

t
u
r
b
u
l
e
n
t

a
n
d

m
o
l
e
c
u
l
a
r

c
o
n
d
u
c
t
i
o
n

a
n
d

b
y

r
a
d
i
a
t
i
o
n

Figure 12: Schematic representation of combustion in a rocket motor; A) hypergolic mixture, B) non-hypergolic mixture
Non-Hypergolic propellants
Homogeneous combustion
Vaporization
Combustion of droplets
Gaseous phase reactions,
diffusion flame
Heterogeneous mixing of the
liquid and gaseous phases
Atomisation and possible mixing
in liquid phase
Heterogeneous
combustion
Reactions in gaseous phase,
premixed flame
Combustion products
D
i
f
f
u
s
i
o
n

o
f

t
h
e

i
n
t
e
r
m
e
d
i
a
t
e

r
e
a
c
t
i
o
n

p
r
o
d
u
c
t
s

i
n

t
h
e

d
i
r
e
c
t
i
o
n

o
p
p
o
s
i
t
e

t
o

t
h
e

f
l
o
w

H
e
a
t

t
r
a
n
s
f
e
r

t
o

t
h
e

l
i
q
u
i
d

a
n
d

g
a
s
e
o
u
s

p
h
a
s
e
s

b
y

t
u
r
b
u
l
e
n
t

a
n
d

m
o
l
e
c
u
l
a
r

c
o
n
d
u
c
t
i
o
n

a
n
d

b
y

r
a
d
i
a
t
i
o
n

Secondary atomization
Gaseous phase mixing
The combustion process for non-hypergolic propellants is somewhat similar. The
possible reaction phenomena are shown in Figure 12b combustion takes place either
in a heterogeneous mixture of a liquid or gaseous phase or in a homogeneous mixture
of atomised propellants. In the first case droplets of one of the propellants will react
with the surrounding gas of the other propellant in a diffusion flame. In the second
case, the homogeneous mixture of gases reacts with a premixed flame.
The time available for the flow to vaporize, mix, etcetera is called the dwell time or
residence time and can be expressed as:
c c
/U L = t
(4-1)
L
c
is length of combustor and U
c
is (average) flow velocity in combustor. Multiplying
denominator and numerator with the gas density in the chamber,
c
, and the chamber
cross-sectional area, A
c
, we get:
( ) /m V A U / A L
c c c c c c c c
= = t
(4-2)
Here V
c
is chamber volume and m is mass flow rate.
Using an earlier derived expression for the characteristic velocity, we get:
* c
* L 1
A
V
* c
1 1
A
V
* c
T R
1
* c
A p
V
2
t
c
2
t
c
c t c
c c

I
=
I
=


= t
(4-3)
Here we introduce the characteristic length, L*, which is defined as the ratio between
the chamber volume and the throat area, A
t
:
t c
/A V L* =
(4-4)
L* is a constant that depends on the type of propellant. Typical values for L* can be
obtained from the next table.
Table 2: L* values for specific propellants
For gaseous oxygen/hydrocarbon fuels, an L* of 1,25 to 2,5 m is appropriate.
For liquid rocket propellants:
Huzel and Huang (1992):
- oxygen-kerosen
e: 1,02 < L* < 1,25 m
- oxygen-hydrogen: 0,76 < L* < 1,02 m
- oxygen-hydrogen (hydrogen is injected as gas): 0,56 < L* < 0,71 m
- nitrogen tetroxide/hydrazine based fuel: 0,60 < L* < 0,89 m
- hydrogen-peroxide/RP-1: 1,52 < L* < 1,78 m (including catalyst bed)
Barrre et al. (1960):
- oxygen- ethyl alcohol: 2,5 < L* < 3 m
- nitric acid - UDMH: 1,5 m < L* < 2,5 m
- nitric acid - hydrocarbons: 2,0 m < L* < 3,0 m
and Dadieu, Damm and Schmidt:
- LOX - gasoline: 1,5 m < L* < 2,5 m
- nitric acid - UDMH: 1,5 m < L* < 2,0 m
- nitromethane (monopropellant): L* = 4,0 m (including catalyst bed)
This shows that the minimum chamber size that still guarantees a satisfactory
combustion process, is found from the minimum value of L*. The residence time
depends most on the slowest process taking place in the chamber; this is generally
the vaporization of the propellants.
As far as large-sized combustion chambers are concerned, it is possible to detect a
trend in the evolution of the shape, see Figure 11. When the first chambers were
developed, the design of the injection system and nature of the propellants were such
that, in order to obtain a satisfactory combustion quality, it was necessary to employ
long lengths. The chamber cross section was large compared to the throat area to
ensure low heat transfer to the wall (material resistance was low). The progress made
in injection system, in cooling system, and the development of new materials and new
propellants made it possible to reduce the length of the chamber and to reduce the
cross-section of the chamber. Further improvement of these parameters led to a
further reduction of size of the chamber, while the nozzle has increased in size.
4.3 Combustion stability
4
NASA specifications allow up to 5% of the chamber pressure oscillations for stable
combustion (Huzel et al 1992). Therefore, if only 5% of the combustion pressure was
selected as the pressure drop through the injector, and a local pressure disturbance of
5% of the combustion pressure occurred, the flow through the injector would stop. This
would then cause an increase in pressure, and therefore a temporary rise in the local
flow rate. If this consecutive drop and rise in pressures occurs close to any of the
systems natural frequencies, combustion instabilities will develop. Hence, the injector
pressure drop must be sufficient to provide isolation between a combustion
disturbance and the local, instantaneous propellant flow rates.
Injector pressure drop requirement differs with the type of injector considered.
According to [Humble et al], the pressure drop requirement is 10-15% for unlike-
impinging injectors and 20-25% for like impinging injectors. For concentric injectors,
the pressure drop requirement may be
as low as 5%.
4.4 Pressure drop due to flow acceleration
Fluid entering the chamber via the injector will vaporize and react forming a low
density gas mixture. Because of conservation of mass, this will lead to an increase in
flow velocity. When the flow is turbulent, the velocity distribution across the chamber is
relatively uniform, so that the longitudinal momentum of the flow is approximately
equal to the product of mass flow and flow velocity. The change in longitudinal
momentum must be balanced by the pressure difference applied between the two
ends of the passage:
( ) ( )
2 2
c c c c c i
c c i c i
M p v p p
v m v v m p p A
dv m dp A
= =
~ =
=

(4-5)
Here p
i
is the pressure just behind the injector face, and p
c
, v
c
, and M
c
respectively are
the pressure, velocity and Mach number at the end of the combustion chamber just in
front of the convergent section.
The pressure drop occurring in the combustion chamber as a function of the Mach
number is presented in the next figure.

4
Instability in the combustion seems, at best, to increase the local heat transfer rates, which
often leads to burning of injector plates or other walls, and at worst it may cause oscillations in
pressure large enough to lead to explosions. In current testing practice any detectable vibration
is generally the signal
Figure 13: Pressure drop in combustion chamber
The Figure 13 clearly shows that to limit the pressure drop in the chamber, it is
important to keep the flow velocity as low as possible.
To reduce losses due to flow velocity of gases within the chamber, the combustion
chamber cross sectional area should be at least three times the nozzle throat area.
This limits the flow velocity in the combustion chamber to a maximum of about M =
0.3. In practical cases, the contraction ratio mostly is taken larger than 3. For example
for the HM-60 a value of 5 is used. Using the earlier determined HM-60 combustor
volume, it follows for the length of the combustor a value of less than about ~ 0.2 m.
4.5 Catalyst bed
Monopropellant decomposition chambers contain a catalyst bed with a catalyst to
further the decomposition of the propellant. This catalyst bed is contained by retainer
gauzes, see Figure 14. The propellant flow is spread evenly over the catalyst bed by
the injector. Sometimes the retainer gauzes assist in the spreading of the
monopropellant over the bed. They also support the catalyst bed to prevent
deformation.
Important for the design of the catalyst bed is a
large surface area in a small volume. Most
hydrazine thrusters use a catalyst bed made
from iridium impregnated alumina pellets 1.5 to
3 mm in diameter (smallest pellets in upper
catalyst bed). An alternative is to use finely
divided iridium on an aluminum oxide support
or platinum-iridium mesh. A typical catalyst bed
for a 1 N hydrazine thruster is 25 -50 mm long
and 6.5 mm in diameter for a hydrazine loading
of 0.015-0.060 gram/mm
2
-s. Generally iridium
is present to the extent of 30% of the total
catalyst mass. To allow proper operation of the
catalyst, the catalyst bed may be conditioned
by one or more heater elements.
In case of hydrogen peroxide propellant, silver
wire cloth and/or silver plated nickel screens are used as catalyst. Typical reactivity
data for hydrogen peroxide with silver catalyst have been determined by [Bengtson]
and are 9.4 to 11.4 g of 85% H
2
O
2
/(minute-cm
2
) using silver plated nickel screens of
28 x 28 mesh, with a wire diameter of 0.19 mm. According to [Jonker], attention must
be paid to differences in thermal expansion between the catalyst bed and the chamber
as this might lead to a wet start. The nickel based silver plated screens increase
temperature compatibility compared to silver wire cloth. Due to the melting points of
Figure 14: Monopropellant thruster
both pure silver and brass, the concentration of hydrogen peroxide is kept at a
maximum of 85%.
The life of the catalyst bed may be an important factor in thruster life. This is for
instance the case for hydrazine thrusters, where life of the catalytic bed mainly
depends on the degradation that occurs in the bed [Brown]. This depends on 1)
mechanical failure of the catalyst pellets, and 2) reduction of catalytic activity caused
by impurities on the surface of the pellets. To avoid mechanical failure, the catalyst
bed is preheated to a temperature of 200 -315 degree Celsius. To avoid reduction of
catalytic activity, super-pure hydrazine is used.
The flow through a catalyst bed usually is accompanied by a strong pressure drop.
This in part is associated with the reactions taking place, but also because the catalyst
bed provides for flow blockage which has to be overcome. A possible approach might
be to use relations for the pressure drop that occurs for single phase flow through a
tube filled with a porous medium, see for instance the work of Ergun [Levenspiel] and
than add a correction for two-phase flow. However, further investigation is needed.
4.6 Chamber throat assembly
The connection of the chamber/combustor tube to the nozzle is mostly through a
throat assembly. In most rocket motors this is a convergent-divergent nozzle part,
sometimes integrally connected to the tubular section of the chamber that allows for
the chamber to be tested at sea level conditions (separate from the nozzle) without
flow separation. The design of the convergent is mostly aimed at reducing pressure
losses due to the flow contraction; see sections on liquid injection and nozzle design.
For small combustion chambers the convergent volume is about 1/10th the volume of
the combustor tube.
4.7 Combustor tube wall geometry
Various combustor tube wall geometries can be distinguished. The choice is mostly
governed by the heat loads and the associated cooling required.
Radiation-cooled rocket engines are mostly of a single wall design, where the
structural material is capable of carrying both the heat and pressure loads. Usually a
coating is applied to protect the material from oxidation.
Rocket engines with high heat loads mostly use the double wall design. Most high
performance combustion chambers are of a double wall design allowing efficient
removal of excess heat either through regenerative or dump cooling. Early combustor
(thrust chamber) designs had low chamber pressure, low heat flux and low coolant
pressure requirements, which could be satisfied by a simple "double wall chamber"
design with regenerative and film cooling.
For higher heat flows, "tubular wall" combustion chambers are used. This is by far the
most widely used design approach for the vast majority of large rocket engine
applications. These chamber designs have been successfully used for amongst others
the Ariane 5 HM-60, the Japanese LE5, and the USA H-1, J-2, F-1, and RS-27 rocket
engines.
To cope with still higher heat flow, channel wall"
combustors are used. These are so named because
the coolant flows through rectangular channels, which
are machined or formed into a hot gas liner fabricated
from a high-conductivity material, see figure 14. The
figure shows that the wall consists of three layers: a
coating, the slotted high-conductivity material, and the
close-up. These three layers can be different materials
or the same.
Figure 15: Example of channel wall
4.8 Chamber materials
Chamber materials used are primarily selected based on their ability to withstand the
combined heat and pressure load as well as their compatibility with the coolant fluid. In
addition, for multi-shot (pulsed) engines/thrusters, the resistance to fatigue as well as
stress corrosion is important.
Typical materials used for high-thrust liquid rocket engines is stainless steel as
structural material and as high-conductivity material some kind of copper or nickel
alloy to transfer the heat to a coolant. A typical such copper alloy is NARloy Z with a
thermal conductivity of 330 W/m-K. Some rocket motors also use a refractory metal
like Niobium as the structural material.
Bi-propellant RCS thrusters and resistojets mostly use refractory metals like rhenium,
molybdenum, columbium (Niobium) and alloys of these elements as the structural
material. Advantage of these materials is that they can withstand very high
temperatures. However, they are very susceptible to oxidation. To this end, usually a
silicide coating is used to provide protection against the aggressive combustion gases.
Recently, one is considering the use of ceramic-matrix carbon as the structural
material as this requires no coating and is equally capable of attaining high
temperatures. As the heat loading of the injector is less than for the combustor tube
and contraction, titanium can be used as the structural material.
Hydrazine monopropellant thrusters typically experience decomposition temperatures
in the range 1000-1500 K at a chamber pressure up to about 20 bars. Typical
structural material for hydrazine monopropellant thrusters is either stainless steel or
brass (copper alloy C3600). Sometimes also nickel alloys as Haynes 25 or 188 are
used.
Thrust chambers for hydrogen peroxide monopropellant thrusters can be fabricated
from either stainless steel or brass (e.g. copper alloy C36000). The thermal expansion
rate of the latter closely matches that of the silver catalyst, which makes that as the
chamber heats up both the chamber and catalyst expand at a similar rate. Brass also
has the advantages of being easily machined and of having high strength.
Choice of chamber material depends on the use of the material (structural material,
insulation, and conductor) and considerations concerning strength, density, corrosion
resistance
5
, fatigue resistance, brittleness, etc. For an explanation of these terms, you
are referred to for instance material handbooks.
Material properties for a range of structural materials used in rocket design have been
collected in [SSE].
4.9 Chamber wall thickness based on internal pressure
Once chamber shape and volume are determined, we can determine the chamber
wall thickness. This thickness greatly depends on the internal pressure. Typical values
used range from a few bar for small spacecraft engines up to 200 bar for the Space
Shuttle Main Engine. Besides internal pressure several other loads exist that should
be considered when determining the wall thickness. Typical such loads are e.g.
handling loads or because of thermal gradients. It also may be that thickness is limited
from manufacturing; minimum thickness is about 0.1-0.2 mm for stainless steel, 0.2
mm for aluminium, and 0.5 mm for titanium. In this section, we consider only the effect
of internal pressure loading.

5
Material compatibility with a propellant is classified sequentially from Class 1 materials, which
exhibit virtually no reaction with the propellant, to Class 4 materials, which react strongly with the
propellant.
To estimate chamber wall thickness, we use thin shell
6
theory [Megson]. In case of a
fully metallic spherical chamber, the wall thickness simply follows from the relation for
the circumferential (hoop) stresses existing in the walls due to this loading:
o

=
2
r p
t (4-6)
Where o is ultimate or yield strength of material, t is thickness of shell, p is internal
pressure, and r is radius of tank.
According to [Megson], shell thickness for a cylindrical section is twice that of a
sphere. Next to internal pressure loads also other loads, like heat loads and handling
loads, do influence wall thickness. Furthermore, also a minimum thickness may be
required to allow for welding, etc. For more details, see chapter Design of thin shell
structures.
An important parameter in the comparison of materials is the specific strength. This is defined as
the ratio between strength and density. The higher the specific strength the stronger or the lighter
the structure will be. Sometimes specific strength is expressed in m
2
s
2
. For example for titanium,
this gives a value of 23 x 10
4
m
2
s
2
.
4.10 Chamber mass
Chamber mass can be estimated based on shell mass:
t S M
shell
=
(4-7)
With is density of shell material, S is shell surface area, and t is wall thickness. In
case the shell consists of multiple layers, like when dealing with composite over-
wrapped chamber walls, we get:
n
n
n n shell
t S M =
_
(4-8)
Where n refers to the various material layers.
Chamber mass follows from:
shell chamber
M K M =
(4-9)
Where K is correction factor taking into account additional mass items like mounting
provisions, thermal insulation, and provisions for cooling.
More information can be found in the chapter entitled Thrust chamber mass.
4.11 Chamber service life
A well known phenomenon in the field of structural engineering is that repeated
stressing of a material can cause failure, even when the applied stress is well below
the yield stress. This is referred to as low cycle
7
thermal (due to differences in
expansion) fatigue. According to Sutton, low cycle fatigue is one of the most important

6
Thin shell theory can be applied in case shell thickness is limited to less than 10% of the radius
of curvature of the shell.
7
Less than 1000 cycles.
causes of rocket combustion chamber failing with cracks sometimes appearing
already after the first burn.
For more information, see the chapter entitled Design of thin shell structures.
4.12 Other chamber characteristics
Besides mass, and size also other parameters, like cost, reliability, and safety, are of
importance to consider when designing a rocket thrust chamber.
Chamber production and development costs depend on chamber type, shape, size,
and lot acceptance testing. Production costs furthermore depend on quantity.
Typical thruster reliability data show a failure rate of 5.7 x 10
-8
failures per hour. This
gives a reliability of 0.995 over a 10 year life.
To estimate these characteristics for conceptual design purposes, it is advised to use
either estimation by analogy of parametric estimation. For later stages, we can use
either parametric or engineering build up estimation. It is for the reasons of estimation
that it is advised to develop a data base with actual (historic) values on the
characteristics of interest and to the level of detail considered necessary.
Problems
1. You are designing a 407 N (vacuum thrust) rocket motor using MMH and NTO as
propellants. Mass mixture ratio is 1.65, which gives a flame temperature of 3056.8
K and a vacuum specific impulse of 314 s. Thruster inlet total pressure is 15.2 bar.
Calculate for this rocket motor for a chamber pressure of 10.9 bar:
a) Mass flow
b) chamber length (including convergent part) and diameter
c) throat diameter,
d) injection velocity,
e) area of one single injector hole (both for MMH and NTO), and
f) total number of injection holes
Discuss the effect of halving thruster inlet total pressure.
For the calculation of the number of injection holes you may come up with your
own distribution of the injection holes over the injector (motivate).
References
1. Barrre M., Jaumotte A., Fraeijs de Veubeke F., and Vandenkerckhove J., Rocket
Propulsion, Elsevier Publishing Company, 1960.
2. Bejan A, Heat Transfer, John Wiley and Sons Inc., New York, 1993.
3. Bengtson E., website http://www.peroxidepropulsion.com, June 2005.
4. Brown C.D., Spacecraft propulsion, AIAA Education Series, 1995.
5. Huzel D.K. and Huang D.H., Design of liquid-propellant rocket engines, NASA
SP-126, 1971.
6. Humble R, et al., Space Propulsion Analysis and Design, Space Technology
Series, McGraw-Hill Companies Inc. 1995.
7. Jonker W., Mayer A.E.H.J., and Zandbergen, B.T.C., Development of a rocket
engine igniter using the catalytic decomposition of hydrogenperoxide, 8
th
International hydrogen peroxide propulsion conference, Purdue University, West
Lafayette, Indiana USA, September 2005.
8. Levenspiel O., Engineering Flow and heat Exchange, Revised ed., Plenum 1998.
9. Megson T.H.G. Aircraft Structures for engineering students, Edward Arnold, 3rd.
Ed. ISBN 03407-05884.
10. PBNA Polytechnisch Zakboekje, 48
th
edition, 1998.
11. SSE, Propulsion web pages.
12. Sutton G.P., Rocket Propulsion Elements, 6
th
edition, John Wiley & Sons Inc.,
1992.
For further reading
1. Thrust chamber life prediction (NASA-CR-134806, 1975)

S-ar putea să vă placă și